You are on page 1of 11

Article https://doi.org/10.

1038/s41467-023-39192-z

Resolving nanostructure and chemistry of


solid-electrolyte interphase on lithium
anodes by depth-sensitive plasmon-
enhanced Raman spectroscopy

Received: 22 December 2022 Yu Gu 1,5, En-Ming You 1,5, Jian-De Lin1, Jun-Hao Wang1, Si-Heng Luo1,
Ru-Yu Zhou1, Chen-Jie Zhang2, Jian-Lin Yao2, Hui-Yang Li1, Gen Li 1,
Accepted: 31 May 2023
Wei-Wei Wang1, Yu Qiao 1, Jia-Wei Yan 1, De-Yin Wu 1, Guo-Kun Liu3,
1234567890():,;
1234567890():,;

Li Zhang1, Jian-Feng Li 1, Rong Xu4, Zhong-Qun Tian 1 , Yi Cui 4 &


Bing-Wei Mao1
Check for updates

The solid-electrolyte interphase (SEI) plays crucial roles for the reversible
operation of lithium metal batteries. However, fundamental understanding of
the mechanisms of SEI formation and evolution is still limited. Herein, we
develop a depth-sensitive plasmon-enhanced Raman spectroscopy (DS-PERS)
method to enable in-situ and nondestructive characterization of the nanos-
tructure and chemistry of SEI, based on synergistic enhancements of localized
surface plasmons from nanostructured Cu, shell-isolated Au nanoparticles and
Li deposits at different depths. We monitor the sequential formation of SEI in
both ether-based and carbonate-based dual-salt electrolytes on a Cu current
collector and then on freshly deposited Li, with dramatic chemical recon-
struction. The molecular-level insights from the DS-PERS study unravel the
profound influences of Li in modifying SEI formation and in turn the roles of
SEI in regulating the Li-ion desolvation and the subsequent Li deposition at SEI-
coupled interfaces. Last, we develop a cycling protocol that promotes a
favorable direct SEI formation route, which significantly enhances the per-
formance of anode-free Li metal batteries.

Lithium metal is regarded as an ideal anode for rechargeable Li batteries Li deposition/dissolution and thus the reversible operation of Li metal
because of its ultrahigh specific capacity1–3. However, the electro- batteries7,8. However, fundamental understanding of the SEI formation
chemical plating/stripping of Li metal usually presents a dendritic and evolution during battery operation is still limited, which limits the
morphology, leading to low Coulombic efficiency and poor cycling development of high-performance Li metal batteries.
stability of the Li metal anode4–6. The solid-electrolyte interphase (SEI), a The SEIs formed during the battery operation usually present
nanometer-thin layer formed at the electrode-electrolyte interface due structural and compositional heterogeneity – nanometric phases
to the electrolyte decomposition, plays a crucial role in determining the of decomposition products from the electrolyte are dynamically

1
State Key Laboratory of Physical Chemistry of Solid Surfaces, iChEM, Innovation Laboratory for Sciences and Technologies of Energy Materials of Fujian
Province (IKKEM), College of Chemistry and Chemical Engineering, Xiamen University, Xiamen, China. 2College of Chemistry, Chemical Engineering and
Materials Science, Soochow University, Suzhou, China. 3State Key Laboratory of Marine Environmental Science, College of the Environment and Ecology,
Xiamen University, Xiamen, China. 4Department of Materials Science and Engineering, Stanford University, Stanford, CA, USA. 5These authors contributed
equally: Yu Gu, En-Ming You. e-mail: zqtian@xmu.edu.cn; yicui@stanford.edu; bwmao@xmu.edu.cn

Nature Communications | (2023)14:3536 1


Article https://doi.org/10.1038/s41467-023-39192-z

precipitated and distributed heterogeneously in the SEI7,9. The for- SHINERS to in-situ characterize the dynamic processes of SEI forma-
mation and evolution of SEIs become more elusive when the anode- tion and evolution during the operation of Li metal batteries. The
free configuration of Li metal batteries is employed10–13, in which SEI synergistic plasmonic enhancement of nanostructured Cu, shell-
with different components and structures needs to form sequentially isolated Au nanoparticles, and Li deposits enables a depth-sensitive
on the surface of the Cu current collector and then on the deposited detection of signals from the SEI with a thickness of tens of nanometers
Li14–16. Although SEIs have been intensively investigated by various and its interfaces to electrode (Cu or Li) and electrolyte (Fig. 1a). We
advanced experimental techniques17–26 and theoretical simulations27,28, further elucidate the formation of SEIs through two different routes,
the preceding SEI formation on the Cu in the anode-free configuration the sequential formation and the direct formation, highly depending
and its possible detrimental effects on Li deposition-dissolution on the cycling protocol. This comprehensive molecular-level infor-
cycling stability is often overlooked. Therefore, in-situ methods that mation from the DS-PERS enriches our understanding of SEI formation
can follow dynamic interfacial processes and provide accurate and evolution during Li plating. Based on this understanding, we fur-
molecular-level information is urgently needed to gain more detailed ther design the favorable SEIs for practical anode-free Li metal using a
insights into the complicated interfacial processes involving SEI for- potentiostatic-galvanostatic polarization strategy. The electrodes
mation and evolution and to guide the design of favorable SEIs to engineered with this type of SEI exhibited significantly enhanced
improve the performance of Li metal batteries. cycling stability and an expanded lifetime in the anode-free Li metal
Surface-enhanced Raman spectroscopy (SERS) is a surface- batteries. The DS-PERS method provides tremendous opportunities
sensitive technique based on the effect of localized surface plasmons for nondestructively characterizing the nanoscale interphases/inter-
(LSPs) of especially free-electron metal (e.g., Au, Ag, Cu and Li) faces with spatial information, which is a great challenge not only in the
nanostructures29–31. It is able to characterize the interfacial processes battery field but also in the general fields of material sciences and
with molecular-level fingerprint information over a wide spectral energy sciences.
range32–35, thereby showing great potential to identify the composition
and structure of SEIs in batteries36–43. However, since the LSP Results and Discussion
enhancement exhibit an exponential decay of electromagnetic inten- Working principle of DS-PERS
sity within a couple of nanometres32, its application in characterizing The DS-PERS is based on a SERS-SHINERS integrated plasmonic
the SEI with a thickness of tens of nanometers is limited – only the enhancement structure with unique depth sensitivity by employing a
information from the vicinity of hotspots with the large local electro- SERS-active nanostructured Cu substrate, and shell-isolated nano-
magnetic field can be extracted by SERS (see Supplementary Note 1). particles (SHINs) introduced initially on top of the Cu substrate sur-
Alternatively, shell-isolated nanoparticle-enhanced Raman spectro- face, as illustrated in Fig. 1a and Supplementary Notes 2–5. The SHINs
scopy (SHINERS) is developed, in which plasmonic metal cores are had plasmonic Au cores (~60 nm) with ultrathin and pinhole-free SiO2
coated with ultrathin inert shells to separate the cores from the sub- shells (~2 nm), and the nanostructured Cu was created by the elec-
strate and are used solely as Raman signal amplifiers, giving reliable trochemical oxidation-reduction roughening method54 (Supplemen-
signals from the sample surface44–51. Monitoring of dynamic nature of tary Figs. 1, 2). Before Li deposition, the junctions of adjacent
SEI formation on anodes in Li-ion batteries via SHINERS has been nanostructured Cu serve as hotspots for providing the finger-print
reported, including a recent study involving Si anodes52. However, information of SEI formed in the vicinity of Cu surface (i.e., the inner
using the SHINERS to characterize the SEIs in Li metal batteries has region of SEI) in the early stage of formation (Fig. 1b). The LSPs-active
rarely been reported53, mainly due to the complex operation of SHI- SHINs provide additional signals from the hotspots of the SHINs and
NERS at the highly reactive Li metal surfaces and in organic electrolytes. Cu such that the chemical and structural information of SEIs at outer
In this work, we develop a depth-sensitive plasmon-enhanced regions can be obtained (Fig. 1c). After Li deposition, the Li metal with
Raman spectroscopy (DS-PERS) method that combines the SERS and SHINs generated new hotspots (Fig. 1d) to further enhance Raman

a b SEI formation c SEI further growth d SEI restructuring


Electrolyte Laser
Raman
hotspot signals

III' III'
III III III III III III SEI
SHIN SEI
II' II'
II II SEI II II II II Li II' II'
II II II II II II I'
I I I

Cu Cu Cu Cu

e
Electrolyte
SiO2 shell Outer signals (by II & III) Outer signals (by II' & III')
Inner signals
Au Au + +
core core (excited by I & II) Inner signals (by I & II) Inner signals (by I' & II')
SEI Au
core
Li Li Outer layer Outer layer
Cu Cu Inner layer
Li Li Inner layer Inner layer
Li
0 10
log(Eloc/E0)4 Cu Cu Cu

Fig. 1 | DS-PERS by synergistic LSP enhancement pertinent to SEIs. from the SEI. e The distribution of simulated electromagnetic field near the inte-
a–d Schematics of the DS-PERS that enables detection of the signals from different grated plasmonic substrates during Li deposition. E1oc and E0 represent the loca-
depths in the SEI. The plasmonic structure of nanostructured Cu, SHINs and lized field and the incident field, respectively. Scale bar: 60 nm.
deposited Li generates strong electromagnetic field to enhance the Raman signals

Nature Communications | (2023)14:3536 2


Article https://doi.org/10.1038/s41467-023-39192-z

a 30 cps

Bulk electrolyte
b 30 cps

Bulk electrolyte


×0.025 ×0.025

Cu Cu + SHINs
~400 ~540
~1340 ~1540
1.2 V I+II 1.2 V
~400 ~540 ~1540 ~1190
Intensity (a.u.)

Intensity (a.u.)
I 0.8 V II+III 0.8 V
~700

I 0.2 V III 0.2 V

Cu@Li ~1050 Cu@Li + SHINs


~1050
~640
~380 1850
~920
I' 0.1 V I'+II' 0.1 V
~530
I' 0.2 V II'+III' 0.2 V

300 600 900 1200 1500 1800 300 600 900 1200 1500 1800
Raman shift (cm1) Raman shift (cm1)

c
Depth-fixed detection vs. Depth-sensitive detection
Cu-SEI Li-SEI Cu-SEI Li-SEI
Outer layer III III III' III'
Inner layer SEI restructuration
Li II II Li II' II'
Li II II II' II' Li
plating I' plating I'
I I
Cu Cu Cu Cu

ROLi, Li2NSO2CF3, Li3N LiOH, LiF, Li2SxOy, Li2O, Li2CO3 Li2SxOy, LiF, Li3N, ROLi LiF, Li2S, Li2O, Li3N

Fig. 2 | Characterizations of SEI formed in the 1 M LiTFSI in DME-DOL using DS- sensitive strategy based on synergetic LSPs of integrated Cu-SHINs for in-situ
PERS. a, b In-situ Raman spectra showing the sequential formation and evolution of probing the formation and restructuring of SEI. By the DS-PERS, it is revealed that
SEIs on the nanostructured Cu substrate (a) and the Cu-SHINs substrates (b) before the primary Cu-SEI can undergo restructuring with the participation of freshly
and after Li deposition. Raman spectra of the bulk electrolyte are displayed for deposited metallic Li to form low-oxidation-state species with highly amorphous
comparison. Peaks marked by asterisks are from the electrolyte. c Schematic and heterogeneous structure.
comparison of the depth-fixed strategy based on LSPs of Cu alone and the depth-

signals of SEI near the Li surface and in the outer region by plasmonic 1010 (Supplementary Figs. 5, 6 and Supplementary Table 1). Second,
coupling. The key advantage of such an elaborative design is the two SHINs locations, on top of and within the Li deposit, were
synergistic plasmonic enhancement effect, which provides higher observed upon Li deposition. Only the SHINs on the Li deposits could
detection sensitivity and depth sensitivity of Raman signals from SEIs generate effective hotspots to maintain a high EF of up to 1010
and related interfaces. In this circumstance, we can capture the depth- (Supplementary Note 5 and Supplementary Fig. 7). The above pro-
dependent nanostructure and chemistry of SEI during its formation posed plasmonic enhancement strategy with multiple hotspots
and evolution at different stages of Li plating. provides a basis for the sensitive detection of Raman signals of the
The synergetic LSP enhancement was modeled using the finite SEI during its formation.
element method (Fig. 1e and Supplementary Figs. 2–4 and Supple- To validate the synergetic LSP enhancement, we first performed
mentary Table 1). For the Cu-SHINs integrated substrate (Supple- the in-situ Raman measurements on the Cu electrode in an ether-based
mentary Note 3 and Supplementary Fig. 3), the hotspots were formed electrolyte of 1 M lithium bis(trifluoromethanesulfonyl)imide in 1,3-
in three types of junctions, namely, Cu island-to-Cu island (type-I), Cu dioxolane and 1,2-dimethoxyethane (LiTFSI in DME-DOL)1,55. The Cu
island-to-SHIN (type-II) and SHIN-to-SHIN (type-III) junctions. The working electrodes with and without SHINs were assembled in a sealed
enhancement factor (EF) of such a configuration can reach up to 1010 three-electrode Raman cell where Li foils served as both the reference
(Supplementary Table 1), sufficiently high for the ultrasensitive and counter electrodes (Supplementary Fig. 8a). A thin liquid layer
detection of the Raman signals from the SEI. However, the distribu- (< 100 μm) was trapped between the Cu surface and quartz window to
tion of SHINs on the Cu surface changed after primary SEI formation ensure that the Raman signals were from the metal-SEI interface,
and Li deposition (Supplementary Fig. 4), and we need to consider instead of the bulk electrolyte (Supplementary Fig. 8b). This was
how the change of SHINs locations influenced the EF. First, SEI for- confirmed by the potential-dependent inversion of the intensity ratio
mation was initiated only at the metal surface, leaving SHINs at least of two prominent bands at 743 and 942 cm−1, which are from the
partially embedded in the SEI (Supplementary Note 4). However, vibration modes of the TFSI anion (S–N and C–S stretching, CF3
since SEI components absorb light weakly and their dielectric con- bending) and DOL solvent (C–O and C–C stretching), respectively
stants are close to the electrolyte, the SEI growth will not sub- (Fig. 2a, b, and Supplementary Figs. 11, 12).
stantially decrease the strength of the electromagnetic field from the We then conducted the potential-dependent Raman measure-
plasmonic substrates (Supplementary Note 4). Even when the SHINs ments to investigate the electrolyte reduction on a nanostructured Cu
were entirely embedded in the SEI, the average EF was still as high as substrate (Fig. 2a and Supplementary Fig. 11). In the first cycle from 2.0

Nature Communications | (2023)14:3536 3


Article https://doi.org/10.1038/s41467-023-39192-z

to 0.2 V without Li overpotential deposition (OPD), the cyclic voltam- Such a sequential formation and reconstruction of SEIs are overlooked
mograms (CVs) of the Cu electrode show two successive cathodic in the past research on Li metal batteries.
waves starting from ~1.6 V (Supplementary Fig. 10) due to the solvent
reduction and anion reduction, respectively56. However, the bands at Sequential and direct formation of SEIs in carbonate-based
~400, ~540 and ~1540 cm−1 for the reduction products of LiOH/LiF/ electrolytes
Li2SxOy, ROLi/Li2O and Li2CO3/ROCO2Li (Supplementary Table 2), The DS-PERS method was then used to study the formation and evo-
respectively, were not observed until the potential reached 1.0 V. The lution of SEIs in a complex yet promising carbonate-based dual-salt
spectra remained the same upon a further decrease of potential to electrolyte of lithium difluoro(oxalato)borate and lithium tetra-
0.2 V (Fig. 2a and Supplementary Fig. 11), even with SEI continual fluoroborate in diethyl carbonate and fluoroethylene carbonate
growth on the Cu surface. It means that Cu alone, due to the limited (LiDFOB-LiBF4/DEC-FEC) proposed by Dahn and coworkers58,59. We
detection sensitivity of hotspot I, can be used only to probe the fin- apply a galvanostatic polarization (denoted as the G route) to a Cu | |Li
gerprint information of the inner region of the SEI during its initial half-cell, where an integrated Cu-SHINs plasmonic substrate was used
formation (Fig. 2c). Notably, bands for electrolyte components at the as the working electrode. The resulting potential profiles and time-
metal-electrolyte interface were still present even after the potential dependent Raman spectra were recorded simultaneously in Fig. 3a, b.
decreased to 0 V (Supplementary Fig. 11), likely due to a small number As shown in Fig. 3a (black line), the potential on the Cu electrode
of electrolyte molecules buried in the SEI, which was recently observed was decreased to around 0 V within the first 100 s, during which the
by Cui and coworkers57. electrolyte was reduced to the primary SEI on Cu (denoted as Cu-SEI)
When the SHINs were integrated into the nanostructured without the participation of Li OPD. In the corresponding Raman
Cu substrate (Fig. 2b and Supplementary Fig. 12), the Raman bands spectra (Fig. 3b and Supplementary Fig. 18), broad background-like
associated with the electrolyte reduction appeared as early as 1.4 V features in the regions of 300–600 cm−1 and 1200–1600 cm−1 gradually
(Supplementary Fig. 12), indicating the higher detection sensitivity of appeared due to the B-containing decomposition products of LiDFOB
the Cu-SHINs substrate with multiple hotspots (hotspots I and II). salt and the carbonyl species from the reduction of DEC-FEC solvents
When the potential is lower than 1.2 V, new bands at 1340, 1191, and (Supplementary Table 4). At ~90 s, a new band at ~1011 cm−1 appeared
700 cm−1 (excited by hotspots II and III) appeared sequentially (Fig. 2b from the B–O/C–O stretching modes of B-containing components
and Supplementary Fig. 12), which were attributed to the formation of (Supplementary Fig. 18), implying that the primary SEI formed on the
ROLi/Li2SxOy, Li2NSO2CF3/Li2SxOy and Li3N, respectively, from the Cu surface was organic-rich.
successive reductions of TFSI and DOL (Supplementary Table 2). It As the polarization continued, the potential of the Cu electrode
verified that the synergistic LSPR enhancement from the integrated became slightly lower than 0 V, implying the occurrence of Li OPD
Cu-SHINs substrate enabled Raman signals to be detected from dif- (Fig. 3a, red line). Accordingly, dramatic changes in the Raman spectra
ferent depths of the SEI during its formation. Since the SHINs were were observed: bands initially seen when the potential is above 0 V
initially located on the top of the Cu surface and later presented in the disappeared, while four new broad bands associated with organic
outer region of the growing SEI, this observation serves as convincing B-containing species (at ~370 and ~540 cm−1) and Li carbonates/alk-
evidence for the outer surface growth mechanism of SEI (Fig. 2c). oxides (at ~1300 and ~1490 cm−1) appeared (Fig. 3b, Supplementary
Pronounced changes in Raman spectra were observed on both the Fig. 18 and Supplementary Table 4). These species were distributed
nanostructured Cu and integrated Cu-SHINs substrates during a successively from the inner to outer regions of the SEI, as evidenced by
negative potential excursion to −0.1 V where the Li OPD took place comparative Raman spectra recorded on the Cu-SHINs and bare
(lower panels in Fig. 2a, b). The Raman bands of SEI species observed at nanostructured Cu substrates (Supplementary Fig. 19). The intensities
the positive potential disappeared, while strong and broad bands of these bands continued to increase over time until after 30 min,
overlapped in some specific regions for the SEIs formed on the Cu indicating that the SEI formation was completed. Deconvolution ana-
substrate and over the entire wavenumber range for the SEIs formed lysis (Supplementary Fig. 23 and Supplementary Table 4) suggested
on the integrated Cu-SHINs substrate. These phenomena suggest that that the final SEI was enriched by a mixture of organic and inorganic
the SEI was undergoing reconstruction with the participation of freshly species (e.g., B–F compounds and Li2CO3) throughout the thickness of
deposited Li (Supplementary Note 6). The largely broadened shape of the SEI. Some of the mixture species were still in high-oxidation states.
the vibrational bands indicates that the SEI on the Li surface is highly We conclude that under the galvanostatic polarization, the SEI was
amorphous and heterogeneous. Upon further Li deposition, significant sequentially formed by electrolyte reduction at above 0 V (prior to Li
changes in the Raman spectra were observed only for the SEI on the OPD), and followed by further electrolyte reduction below 0 V (with Li
Cu-SHINs substrate (lower panels in Fig. 2a, b), again demonstrating OPD). This transition from the primary SEI on the Cu current collector
that the synergistic plasmonic substrates can create significant Raman (i.e., Cu-SEI) to the final SEI on the Li (i.e., Li-SEI) was accompanied by a
signal enhancement and thus enable depth-sensitive detection of SEI. significant chemical reconstruction, as schematically shown in Fig. 3c.
To study the reconstruction of SEI during Li deposition, we con- The composition and structure of the final Li-SEI were closely
ducted the in-situ Raman measurements using precise control of related to the potential of electrolyte reduction. We found that
applied potential at an interval of 5 mV (Supplementary Fig. 13). The prolonging the bias at the potential prior to Li OPD can promote
resulting broad bands were deconvoluted for component analysis the enrichment of high-oxidation-state organic components of the
(Supplementary Fig. 14 and Supplementary Table 3). Prior to Li OPD, final SEI, especially in the outer region, as confirmed by potential-
the SEI formed on the Cu surface displayed a typical organic-inorganic dependent Raman measurements (Supplementary Fig. 20 and Sup-
hybrid structure, in which high- and low-oxidation-state species (e.g., plementary Table 5). However, when the potential was decreased
ROLi, Li2SxOy and LiF) were the major components of the SEI. However, rapidly to below 0 V, the formation of the primary Cu-SEI was sup-
after the Li OPD was initiated, both the inner and outer regions of the pressed, leading to the direct formation of the SEI on the Li. Herein, we
SEI underwent reconstruction with higher-oxidation-state species design an electrochemically engineered strategy with a potentiostatic-
decomposed to lower-oxidation-state species (e.g., Li2O, LiF, Li2S and galvanostatic polarization (P-G route) to artificially construct a favor-
Li3N). The component analysis was supported by XPS profile analysis able SEI. The P-G route initiate with a negative potentiostatic stepping
(Supplementary Fig. 15). We expect that the role of metallic Li is to (−0.1 V for 100 s, blue region in Fig. 3d) to incur the Li OPD and elec-
promote additional chemical reactions to significantly alter the struc- trolyte reduction for a certain time, followed by a galvanostatic plating
ture of the primary SEI on the Cu current collector to the final SEIs on (~0.5 mA cm−2, purple region in Fig. 3d) to stabilized SEI formation.
the deposited Li (Supplementary Note 6 and Supplementary Fig. 16). Broad Raman bands characteristic of Li-SEI appeared immediately

Nature Communications | (2023)14:3536 4


Article https://doi.org/10.1038/s41467-023-39192-z

Fig. 3 | Formation and evolution of SEIs in the dual-salt electrolyte revealed by Raman spectra recorded during the formation and evolution of the SEI under the
DS-PERS. a–c E–t curve of the galvanostatic polarization (denoted as the G route) P-G route (e). Schematics of the structure and chemistry of the SEI formed under
of Li plating on Cu (a). In-situ Raman spectra recorded during the formation and the P-G route (f). g, h Normalized Raman peak area of bands associated with B–O
evolution of the SEI under the G route (b). Peaks marked by asterisks are from the and carbonate species in sequentially formed Li-SEI (g) and directly formed Li-SEI
electrolyte. Schematics of the structure and chemistry of the SEI formed under the (h). i, j Percentages of specified components in the sequentially formed Li-SEI (i)
G route (c). d, e j–t and E–t curves of electrochemically engineered potentiostatic- and directly formed Li-SEI (j), calculated from the XPS spectra.
galvanostatic polarization (denoted as the P-G route) of Li plating on Cu (d). In-situ

upon the cathodic potential stepping due to the absence of Cu-SEI SEI, as evidenced by the XPS depth profile (Supplementary Fig. 25).
(Fig. 3e and Supplementary Figs. 21, 22). During the following galva- Under the P-G route, since the electrolyte experience a large driving
nostatic plating, the intensities of the broad bands increased markedly force for reduction at the potential below 0 V, DFOB anions can cap-
and approached saturation after only ~15 min, indicating faster evolu- ture electrons from Li metal to enable the B–O or B–F bond cleavage;
tion of SEI than the G route. Importantly, the chemistry and structure thus low-oxidation-state inorganic species such as LiF (Supplementary
of the SEI were significantly different from the sequentially formed SEI Fig. 25) became the main components throughout the Li-SEI. More
under the G route (Fig. 3f), as further revealed by the deconvoluted importantly, the sequentially formed Li-SEI contains more organic
spectra in Supplementary Fig. 23 and Supplementary Table 4. species, especially in the outer region occupied by high-oxidation-
The sequential formation of SEIs under the G route and the direct state organic species such as ROCO2Li and ROLi (Fig. 3g, i). However,
formation of SEI under the P-G route were further investigated using the directly formed Li-SEI exhibited a distinctive polymeric-like struc-
XPS characterization and density functional theory (DFT) calculations ture. The vibrations of O–B–O/CH2 at 1295 cm−1 and Li–O/(LiF)n at
(Supplementary Figs. 24–26 and Supplementary Table 4). We found 470 cm−1 from crosslinked oligomeric borates (e.g., –B(OCH2CH2)3–)
that the broad bands appeared in both routes due to the wide became intensified while carbonate species became weakened
distribution of size and microenvironment of the highly amorphous Li- (Fig. 3h, j and Supplementary Fig. 25). Reactions of the electron-
SEI. Nevertheless, the directly formed Li-SEI by the P-G route had a deficient B-containing intermediates from DFOB anion decomposition
notably reduced thickness compared to the sequentially formed Li-SEI with other SEI components (e.g., electron-rich lithium alkoxide or
from the primary Cu-SEI (Supplementary Fig. 24). Moreover, the lithium semicarbonates) could be one of the reasons that leads to the
organic B–F components from the decomposition of LiDFOB were crosslinking (Supplementary Fig. 26). It is generally acknowledged that
present in the sequentially formed SEI; such components were hybri- the SEIs with polymeric-like structures are beneficial to suppressing Li
dized with inorganic species of LiF with varying amounts across the dendrite growth60,61. Such SEIs composed of oligomers incorporated

Nature Communications | (2023)14:3536 5


Article https://doi.org/10.1038/s41467-023-39192-z

Conventional understanding vs. Renewed understanding by DS-PERS

Normal: Sequential formation of SEI on Cu and then Li


SEI on Cu overlooked
(involving restructuring)

SEI Inferior Li-SEI


Li Inferior Cu-SEI Li
Cu Cu Cu Cu Cu

Improvement: Direct formation of superior SEI on Li


SEI on Cu unchsnged
(suppressing inferior Cu-SEI)

SEI Superior Li-SEI


SEI Li w/o Cu-SEI Li
Cu Cu Cu Cu Cu Cu

Fig. 4 | Schematics of conventional and renewed understandings of SEIs for- unchanged after Li deposition. The DS-PERS revealed a sequential formation of SEI
mation and evolution. In the conventional understanding, either the SEI forma- on the Cu and then on the Li, with a significant chemical and structural
tion on Cu is overlooked, or the SEI formed on Cu is considered to remain reconstruction.

c f Bulk electrolyte
a Li+ anion solvent
DFOB (ring breathing)
SSIP CIP
Intensity (a.u.)

AGG-I
Metal / SEI / Electrolyte

(74.3%)
AGG-I
SSIP AGG-II
CIP
AGG-II
(16.4%)
(9.3%)

700 710 720 730 740 750 840 860 880 900 920 940
d g Sequential SEI/electrolyte
Free DEC/FEC
Intensity (a.u.)

AGG-I
Li or Cu

SSIP
CIP AGG-II
(14.1%) (4.8%) ROBFC2O4Li, ROCO2Li
b Outer layer of SEI
SEI/electrolyte interface 700 710 720 730 740 750 840 860 880 900 920 940
e Free DEC/FEC
h Direct SEI/electrolyte
Intensity (a.u.)

AGG-I

III III

SEI II II
II II
I
LiF, B-(O-CH2-CH2)n
Cu Outer layer of SEI
700 710 720 730 740 750 840 860 880 900 920 940
Raman shift (cm 1) Raman shift (cm 1)

Fig. 5 | The desovlation of Li-ions at different SEI/electrolyte interfaces revealed the SEI/electrolyte interface. The hotspots III enhance the Raman signals from the
by DS-PERS. a Schematic for Li electrodeposition on the electrode with the SEI. The SEI/electrolyte interface. c–e Potential-dependent Raman spectra of Li-ion solva-
solvated ions in the electrolyte need to be desolvated at the SEI/electrolyte inter- tion structure at bulk electrolyte (c), sequentially-formed SEI/electrolyte interface
face and then transported through the SEI to the metal/SEI interfaces for reactions. (d) and directly-formed SEI /electrolyte interface (e). f–h Schematics of Li-ion sol-
b Schematic the schematic of the DS-PERS that can in-situ collect the signals from vation structure governed by the chemical structure of SEIs.

with inorganic species are usually mechanically robust and more facile interface for charge transfer reactions (Fig. 5a). To reveal the roles of
for Li-ion transport. The above information demonstrates that the DS- the SEI/electrolyte interface in regulating the desolvation of Li-ion and
PERS with synergetic LSP enhancements is powerful in elucidating the follow-up Li electrodeposition, DS-PERS was utilized to collect the
different SEI formation mechanisms, and such in-depth understanding enhanced Raman signals from the hotspots III at the SEI/electrolyte
can be truly rewarded by designing the favorable SEI to enable the interface (Fig. 5b).
stable cycling of Li metal batteries (Fig. 4). The Raman spectra of bulk electrolyte (LiDFOB-LiBF4 in DEC-FEC)
are shown in Fig. 5c (left panel). The DFOB anion bands at
Desolvation of Li-ion at SEI/electrolyte interface 700 ~ 750 cm−1 can be deconvoluted into three distinctive bands at 711,
The desolvation of Li ions starts at the SEI/electrolyte interface, fol- 730 and 741 cm−1 associated with solvent-separated ion pairs (SSIPs,
lowed by the Li-ion transport through the SEI to reach the metal/SEI non-coordinated DFOB)/contact ion pairs (CIPs, DFOB coordinating to

Nature Communications | (2023)14:3536 6


Article https://doi.org/10.1038/s41467-023-39192-z

one Li-ion), aggregates-I (AGGs-I, DFOB coordinating to two Li-ions) with the sequentially formed SEI, considerable performance degrada-
and AGGs-II (DFOB coordinating to four Li-ions), respectively (Sup- tion was observed. After 20 cycles, the morphology of deposited Li on
plementary Fig. 27). Obviously, the majority of DFOB anions exists as the Cu with sequentially formed SEI was dendritic and porous (Fig. 6k).
AGGs-I with small amounts of SSIPs/CIPs and AGGs-II due to the strong In contrast, the deposited Li on Cu with the directly formed SEI showed
interaction between DFOB and Li+ (Fig. 5f). This is also verified by the relatively densely packed and flat grains even after 50 cycles (Fig. 6l).
spectra of solvents (Fig. 5c, right panel), where Raman bands from In summary, we developed a DS-PERS method for in-situ char-
fewer coordinated solvents to Li+ are observed. acterization of the formation and evolution of SEI in Li metal batteries.
For the interface between the sequentially formed SEI and elec- The synergistic LSP enhancement from nanostructured Cu, SHINs and Li
trolyte (Fig. 5d), the local solvation structure is different from that in deposits was utilized to enable bottom-up nondestructive detection of
the bulk electrolyte, where the number of free DFOB anions reduces to the structure and chemistry of SEIs and related interfaces with depth
form more AGGs through interactions with multiple Li-ions. We expect sensitivity. The dynamic molecular-level information of the SEI revealed
that the outer SEI region consisting of electron-rich species such as that the sequential formation of SEIs on the bare Cu current collector
ROCO2Li and ROBFC2O4Li can repel the negatively charged DFOB and then freshly deposited Li is a universal phenomenon in both ether-
anions, leading to the coexistence of SSIPs/CIPs and AGGs at this based and carbonate-based electrolytes, which is overlooked in the
interface. However, at the interface between the directly formed SEI conventional understanding of SEI formation. Specifically, the primary
and electrolyte (Fig. 5e), the solvation structure is dominated by AGGs- SEI formed on Cu is composed of less stable high-oxidation-state com-
I, as reflected by the increase of the intensity of free solvents (largely ponents and has to undergo chemical restructuring upon subsequent
weakened interaction with Li+). In this circumstance, inorganic LiF and concurrent Li deposition to complete the formation of a final SEI on Li,
crosslinked oligomeric Li–B(OCH2CH2)3–R are the major components which is less desirable for battery performance. Such a sequential SEI
in the outer region. These species have higher interfacial energy that formation route can be altered by applying an electrochemical strategy
may modify the Li+ solvation environment through the interfacial to promote rapid Li deposition and thus suppress the preceding primary
interactions. For example, the LiF tends to bind with C = O and B‒F in SEI formation on Cu, leading to the direct formation of SEIs on Li with
the DFOB anion. DFT calculations show the large binding energies of the enrichment of more stable low-oxidation-state components. This in-
the LiF coordinated with C = O (−0.33 eV) and the LiF coordinated with depth understanding of SEI and related interfaces guided us to design a
both C = O and B‒F (−0.39 eV), leading to the significant increase in the potentiostatic-galvanostatic polarization strategy for achieving an
proportion of AGGs-I that has spare C = O and B‒F coordination bonds improved polymeric-like structured Li-SEI in a dual-salt carbonate-based
to LiF. In summary, the structure and chemistry of SEI, especially in the electrolyte. The electrodes engineered with this type of SEI exhibited
outer region, can dramatically affect the Li+-anion interaction at the significantly enhanced cycling stability and an expanded lifetime in the
SEI/electrolyte interface. The change in the solvation structure of Li- anode-free Li metal batteries. This study expands the capability of DS-
ions is anticipated to have a significant impact on its desolvation PERS to study the complicated interfaces/interphases in Li metal bat-
behavior and consequently alter the Li deposition. It demonstrates teries, which opens up a new perspective in real-time investigation and
again that the DS-PERS can serve as a powerful tool to reveal the understanding of the mechanisms of SEI formation and evolution and
important roles of SEI/electrolyte interfaces in the Li-ion desolvation the roles of SEI in regulating Li-ion desolvation and Li deposition, which
and Li deposition on the Li metal anode. is pivotal in the rational construction of SEIs and related interfaces for
practical batteries including other alkali metal batteries.
Improve battery performance with the designed favorable SEIs
We expect that the directly formed SEI and its interfaces are favorable Methods
for the stable Li deposition/dissolution on the Li metal anode. The Preparation of nanostructured Cu substrate
improved performance of the electrode with such an SEI is demon- The traditional oxidation-reduction cycle (ORC) method was used for
strated in Fig. 6a–d, where the performance of Cu | |Li cells with Cu electrochemical roughening of the surface of the Cu substrate to
current collectors covered by different SEIs are compared. The elec- obtain nanostructures for effective SERS activation54. The ORC treat-
trolyte used here is the LiDFOB-LiBF4 in DEC-FEC. ment was performed in a 0.1 M KCl aqueous solution using a potential
The cells with directly formed Li-SEI exhibited a long cycle life of step procedure described as follows: oxidation at 0.4 V for 5 s followed
over 200 cycles and an average Coulombic efficiency (CE) as high as by reduction at −0.4 V for 5 s, which was repeated for 3 cycles. The
99.5% (Fig. 6a). In addition, the Li deposition overpotential remained obtained nanostructured Cu substrate was rinsed with deionized water
stable without presenting an obvious voltage hysteresis (Fig. 6c, d). In and rapidly dried under vacuum conditions.
contrast, the one with the sequentially formed SEI showed a much
shorter cycle life (Fig. 6a) and dramatically varied voltage hysteresis Synthesis and assembly of Au@SiO2 SHINs
during cycling (Fig. 6b). Moreover, with directly formed Li-SEI, a uni- Au@SiO2 SHINs were synthesized following previous report44. Au
form Li deposit was formed without apparent dendrite and porous nanospheres with a diameter of ~60 nm were first prepared using the
structures (Fig. 6e, f). The cell performance can be further improved by sodium citrate reduction method. Then, 200 mL of HAuCl4 aqueous
the periodic application of a higher current density pulse of optimized solution (0.01 wt.%) was heated to boiling under string. A total of
magnitude during charge-discharge cycling, leading to the rapid 1.6 mL of sodium citrate aqueous solution (1 wt.%) was quickly added
deposition of Li metal to maintain the high quality of directly formed to the above boiling solution and refluxed for 1 h. Then, the sol was
SEI. It can be obviously observed that the cell with in-situ restoring allowed to cool to ambient conditions for the next step. The ultrathin
showed more stable cycling with a further boosted CE of 99.6% SiO2 shell was then coated by the high-temperature silicate hydro-
(Fig. 6g, h). lyzation method. A 400 μL of (3-aminopropyl) trimethoxysilane
We further assessed the potential of engineering the directly (1 mM) was added to 30 mL of the as-prepared Au nanosphere sol
formed SEI to enable the stable cycling of anode-free full-cells. Proof-of- under stirring without heat. After 15 min of stirring at room tempera-
concept anode-free full cells using the commercial LiNi0.5Mn0.3Co0.2O2 ture, a sodium silicate solution with the pH adjusted to approximately
(NMC532) cathode were tested (Fig. 6i, j). For the cell with the directly 10.0 by H2SO4 was added and stirred for another 5 min. Then, the
formed SEI on the Cu current collector, it delivered the initial discharge mixture was transferred to a boiling water bath and stirred for
capacity of ~176 mAh g−1, and approximately 80 cycles with a high and approximately 30 min to coat the pinhole-free SiO2 shell with ca. 2 nm
stable CE were achieved at a 0.2 C charge/0.5 C discharge rates before in thickness. The as-prepared Au@SiO2 nanoparticles were cen-
the capacity fell to 80%, as shown in Supplementary Fig. 28. For the cell trifuged thrice and washed with deionized water for later use.

Nature Communications | (2023)14:3536 7


Article https://doi.org/10.1038/s41467-023-39192-z

Fig. 6 | Electrochemical performances of anode-free electrodes covered by profiles (h) of Li deposition/dissolution on Cu electrode covered by directly formed
different SEIs. a CE of Li deposition/dissolution on Cu electrodes covered by dif- SEI with periodic in-situ restoring under higher current density (deposition at
ferent SEIs (deposition at 0.5 mA cm−2 and dissolution at 1.25 mA cm−2; capacity of 5.0 mA cm−2 and dissolution at 5.0 mA cm−2). i, j Capacity retention (i) and voltage
1.5 mAh cm−2). b The hysteresis of Li deposition/stripping for the Cu electrodes profiles (j) of the coin-type Cu| |NMC532 cells with Cu electrodes covered by dif-
covered by different SEIs. c, d Voltage profiles of the cells with Cu electrodes ferent SEIs (charge at 0.2 C and discharge at 0.5 C). k, l SEM images of plated Li
covered by sequentially formed SEI (c) and directly formed SEI (d). e, f Cross- morphology in the Cu| |NMC532 cells with Cu electrodes covered by sequentially
sectional SEM images of Li deposited on the Cu electrodes covered by sequentially formed SEI (k) and directly formed SEI (l), respectively. Scale bar: 10 μm.
formed SEI (e) and directly formed SEI (f). Scale bar: 10 μm. g, h CE (g) and voltage

For the assembly of Au@SiO2 SHINs on the surface of the respectively. The distribution of Au@SiO2 SHINs on the nanostructured
nanostructured Cu substrate, the above concentrated sol was diluted Cu substrate after Li deposition as well as the morphology of the Li metal
with 0.5 mL of deionized water to make up a stock solution. Two surfaces after cycling were characterized by SEM (Zeiss GeminiSEM 500)
microliters of the stock solution were then drop-cast onto the surface with an air-isolating transfer apparatus to avoid air exposure of
of a freshly nanostructured Cu substrate and rapidly dried under the samples. XPS measurements were performed on an ESCALAB Xi+
vacuum overnight. (Thermo Scientific) spectrometer using monochromatic Al Kα
(1486.7 eV) X-ray source at 15 KV with a beam spot size of 650 μm. The
Characterization of the samples binding energies were referenced to the C 1 s line at 284.6 eV from
The morphologies of the nanostructured Cu surface and Au@SiO2 adventitious carbon. Depth profiling was fulfilled using Ar ion sputtering
SHINs were characterized by scanning electron microscopy (SEM, in the x-y scan mode at ion acceleration of a 2 kV and ion beam current
HITACHI S-4800) and high-resolution transmission electron microscopy of 2 μA over an area of 2 × 2 mm2. The thickness of the SEI was estimated
(HRTEM, F30) coupled with energy dispersive X-ray spectrometry, based on the calibrated sputtering rate of 5 nm per minute for Ta2O5.

Nature Communications | (2023)14:3536 8


Article https://doi.org/10.1038/s41467-023-39192-z

In-situ electrochemical Raman spectroscopy nanoparticles, whose diameters were 57 or 100 nm, were buried in the
In-situ Raman measurements were conducted on a Raman-11 system Cu substrate. The size of the nanogap was 2 nm for the adjacent Cu
(Nanophoton) equipped with a 50× (NA 0.45) objective and a 300 nanoparticles. The Au core (55 nm)-SiO2 shell (2 nm) nanoparticles
grooves/mm grating. A 785 nm laser was used to avoid the fluores- were modelled as core-shell nanoparticles. The minimal mesh size was
cence background from the electrolyte (Supplementary Fig. 9). The 0.1 nm in close proximity to the nanoparticles and gradually became
homemade sealed Raman cell, in which the nanostructured Cu deco- coarser towards the borders of the simulation domain. The refractive
rated by SHINs served as the working electrode and Li foils served as indexes of the electrolyte and the SEI were 1.4 and 1.6, respectively63.
both the reference and counter electrodes, was assembled in an Ar- The permittivity for Au, Cu, Li and SiO2 were taken from references64,65.
filled glovebox (< 0.1 ppm of H2O and O2). The working electrode Density functional theory (DFT) calculations were implemented with
surface faced upwards and towards a quartz glass window of the cell the B3LYP functional66,67 and 6–311 + G(d, p) basis set. The implicit
with 0.5 mm thickness. The laser power at electrode was approxi- solvent model (SMD)68 was used with acetone (a dielectric constant of
mately 0.7 mW μm−2 and the acquisition time was 60 s for each spec- 20.5) as the default solvent to consider the role of solvent effect.
trum. Raman frequencies were calibrated using a Si wafer. To remove The structures of FEC, DEC, LiDFOB and the corresponding reaction
the interference of fluorescence, the Raman spectra were background- process as well as DFOB-Li+ coordination were performed by using
corrected based on a fifth-order polynomial function. For a higher Gaussian 09 program package69.
signal-to-intensity ratio, while preserving the spectroscopic features, a
peak extraction and retention algorithm62 was applied to the Raman Data availability
spectra recorded on bare nanostructured Cu and integrated Cu-SHINs The data that support the findings of this study have been included in
substrates after Li deposition, as shown in Fig. 2. the main text and Supplementary Information. All other relevant data
supporting the findings of this study are available from the corre-
Measurements of electrochemical performances sponding authors upon request.
All electrochemical tests were carried out in a 2016-type coin-cell
configuration and all cells were fabricated in an Ar-filled glovebox, with References
one layer of Celgard 2325 used as a separator and 50 μL electrolyte. 1. Lin, D., Liu, Y. & Cui, Y. Reviving the lithium metal anode for high-
Galvanostatic charge-discharge cycling was performed on LANHE energy batteries. Nat. Nanotechnol. 12, 194–206 (2017).
CT2001A battery testing system. Two types of SEIs were preformed 2. Liu, J. et al. Pathways for practical high-energy long-cycling lithium
on Cu current collectors with Cu | |Li half-cell configurations before metal batteries. Nat. Energy 4, 180–186 (2019).
electrochemical testing. The sequentially formed SEI on Cu (19 mm 3. Cao, Y., Li, M., Lu, J., Liu, J. & Amine, K. Bridging the academic and
diameter) was prepared by normal galvanostatic route with cathodic industrial metrics for next-generation practical batteries. Nat.
polarization at 0.5 mA cm−2 for 5 h followed by stripping to 0.5 V at Nanotechnol. 14, 200–207 (2019).
1.25 mA cm−2. The directly formed SEI on Cu was prepared by 4. Cheng, X.-B., Zhang, R., Zhao, C.-Z. & Zhang, Q. Toward safe lithium
potentiostatic-galvanostatic route, where a potential of −0.1 V was metal anode in rechargeable batteries: A review. Chem. Rev. 117,
applied for 100 s in the potentiostatic period and a constant cathodic 10403–10473 (2017).
current density of 0.5 mA cm−2 was immediately applied for 1 h in 5. Tikekar, M. D., Choudhury, S., Tu, Z. & Archer, L. A. Design principles
the subsequent galvanostatic period. The residual Li after the for electrolytes and interfaces for stable lithium-metal batteries.
potentiostatic-galvanostatic route was removed by electrochemical Nat. Energy 1, 16114 (2016).
dissolution. For Coulombic efficiency tests, Cu | |Li half-cells were 6. Gu, Y. et al. Surface electrochemistry approaches for under-
cycled by depositing 1.5 mAh cm−2 of Li onto the Cu current collector standing and creating smooth solid-electrolyte interphase and
with different preformed SEIs at −0.5 mA cm−2 followed by stripping to lithiophilic interfaces for lithium metal anodes. Curr. Opin. Elec-
0.5 V at 1.25 mA cm−2. The Coulombic efficiency was calculated by trochem. 26, 100671 (2021).
dividing the total stripping capacity by the total deposition capacity. 7. Peled, E. & Menkin, S. Review—SEI: past, present and future.
For periodic in-situ restoring strategy, a higher current density pulse of J. Electrochem. Soc. 164, A1703–A1719 (2017).
5 mA cm−2 was applied to Cu | |Li half-cell every 20 normal cycles. 8. Wu, H., Jia, H., Wang, C., Zhang, J. G. & Xu, W. Recent progress in
For anode-free full cell tests, Cu | |NMC532 cells were assembled using understanding solid electrolyte interphase on lithium metal
a Cu current collector (19 mm diameter) with differently preformed anodes. Adv. Energy Mater. 11, 2003092 (2020).
SEIs as the anode and commercial NMC532 (Easpring) coated on Al foil 9. Peled, E., Golodnitsky, D. & Ardel, G. Advanced model for solid
(13 mm diameter) as the cathode (94.5 wt.% active). The typical loading electrolyte interphase electrodes in liquid and polymer electro-
of NMC532 was approximately 15 mg cm−2. Before the measurements, lytes. J. Electrochem. Soc. 144, L208–L210 (1997).
all the cells were subjected to two formation cycles at C/10 rate for 10. Qian, J. et al. Anode-free rechargeable lithium metal batteries. Adv.
both charge and discharge processes. These cells were cycled between Funct. Mater. 26, 7094–7102 (2016).
3.6 V and 4.5 V at the C/5 rate for charge and the C/2 rate for discharge. 11. Gu, Y. et al. Lithiophilic faceted Cu(100) surfaces: High utilization of
1 C is equal to 170 mA g−1 of active NMC532 materials. For these con- host surface and cavities for lithium metal anodes. Angew. Chem.
ditions the areal capacity was about 2.5 mAh cm−2. All electrolytes used Int. Ed. 58, 3092–3096 (2019).
were 0.6 M LiDFOB-0.6 M LiBF4/DEC-FEC (2:1, V/V, anhydrous > 99%, 12. Gao, X. et al. Thermodynamic understanding of Li-dendrite forma-
DoDoChem), and all cells were tested at 25 °C. Each test was repeated tion. Joule 4, 1864–1879 (2020).
with three coin cells to ensure consistency. 13. Zheng, J. et al. Regulating electrodeposition morphology of lithium:
Towards commercially relevant secondary Li metal batteries.
Theoretical simulation and calculation Chem. Soc. Rev. 49, 2701–2750 (2020).
The simulated electromagnetic field was obtained by commercial 14. Lin, D. et al. Fast galvanic lithium corrosion involving a Kirkendall-
simulation software (COMSOL Multiphysics) based on the finite ele- type mechanism. Nat. Chem. 11, 382–389 (2019).
ment method. A spherical simulation domain, whose diameter was 15. Wang, W.-W. et al. Probing mechanical properties of solid-
1.8 μm, was created, and perfectly matched layers (PMLs) were electrolyte interphases on Li nuclei by in situ AFM. J. Electrochem.
employed to simulate an open boundary. The bottom half of the Soc. 169, 020563 (2022).
simulation domain was set as Cu. The medium over the Cu substrate 16. Wang, W.-W. et al. Formation sequence of solid electrolyte inter-
was set to be the electrolyte or SEI. A quarter of the spherical Cu phases and impacts on lithium deposition and dissolution on

Nature Communications | (2023)14:3536 9


Article https://doi.org/10.1038/s41467-023-39192-z

copper: an in situ atomic force microscopic study. Faraday Discuss. 37. Ha, Y. et al. Probing the evolution of surface chemistry at the
233, 190–205 (2022). silicon–electrolyte interphase via in situ surface-enhanced Raman
17. Bhattacharyya, R. et al. In situ NMR observation of the formation of spectroscopy. J. Phys. Chem. Lett. 11, 286–291 (2020).
metallic lithium microstructures in lithium batteries. Nat. Mater. 9, 38. Li, H. et al. Surface-enhanced Raman scattering study on passivat-
504–510 (2010). ing films of Ag electrodes in lithium batteries. J. Phys. Chem. B 104,
18. Nie, M., Abraham, D. P., Chen, Y., Bose, A. & Lucht, B. L. Silicon solid 8477–8480 (2000).
electrolyte interphase (SEI) of lithium ion battery characterized by 39. Schmitz, R. et al. SEI investigations on copper electrodes after
microscopy and spectroscopy. J. Phys. Chem. C. 117, lithium plating with Raman spectroscopy and mass spectrometry. J.
13403–13412 (2013). Power Sources 233, 110–114 (2013).
19. Li, Y. et al. Atomic structure of sensitive battery materials and 40. Piernas-Muñoz, M. J., Tornheim, A., Trask, S., Zhang, Z. & Bloom, I.
interfaces revealed by cryo–electron microscopy. Science 358, Surface-enhanced Raman spectroscopy (SERS): A powerful tech-
506–510 (2017). nique to study the SEI layer in batteries. Chem. Commun. 57,
20. Liu, T. et al. In situ quantification of interphasial chemistry in Li-ion 2253–2256 (2021).
battery. Nat. Nanotechnol. 14, 50–56 (2019). 41. Tang, S. et al. An electrochemical surface-enhanced Raman spec-
21. Nanda, J. et al. Unraveling the nanoscale heterogeneity of solid troscopic study on nanorod-structured lithium prepared by elec-
electrolyte interphase using tip-enhanced Raman spectroscopy. trodeposition. J. Raman Spectrosc. 47, 1017–1023 (2016).
Joule 3, 2001–2019 (2019). 42. Rey, I., Lassègues, J. C., Baudry, P. & Majastre, H. Study of a lithium
22. Shadike, Z. et al. Identification of LiH and nanocrystalline LiF in the battery by confocal Raman microspectrometry. Electrochim. Acta
solid-electrolyte interphase of lithium metal anodes. Nat. Nano- 43, 1539–1544 (1998).
technol. 16, 549–554 (2021). 43. Naudin, C. et al. Characterization of the lithium surface by infrared
23. Wang, W.-W. et al. Evaluating solid-electrolyte interphases for and Raman spectroscopies. J. Power Sources 124, 518–525 (2003).
lithium and lithium-free anodes from nanoindentation features. 44. Li, J. F. et al. Shell-isolated nanoparticle-enhanced Raman spec-
Chem 6, 2728–2745 (2020). troscopy. Nature 464, 392–395 (2010).
24. Zhou, Y. et al. Real-time mass spectrometric characterization of the 45. Su, M., Dong, J.-C. & Li, J.-F. In-situ Raman spectroscopic study of
solid–electrolyte interphase of a lithium-ion battery. Nat. Nano- electrochemical reactions at single crystal surfaces. J. Electrochem.
technol. 15, 224–230 (2020). 26, 54–60 (2020).
25. Gao, Y. et al. Unraveling the mechanical origin of stable solid 46. Hy, S. et al. In situ surface enhanced Raman spectroscopic studies
electrolyte interphase. Joule 5, 1860–1872 (2021). of solid electrolyte interphase formation in lithium ion battery
26. He, X., Larson, J. M., Bechtel, H. A. & Kostecki, R. In situ infrared electrodes. J. Power Sources 256, 324–328 (2014).
nanospectroscopy of the local processes at the Li/polymer elec- 47. Cabo-Fernandez, L., Bresser, D., Braga, F., Passerini, S. & Hardwick,
trolyte interface. Nat. Commun. 13, 1398 (2022). L. J. In-situ electrochemical SHINERS investigation of SEI compo-
27. Christensen, J. & Newman, J. A mathematical model for the lithium- sition on carbon-coated Zn0.9Fe0.1O anode for lithium-ion batteries.
ion negative electrode solid electrolyte interphase. J. Electrochem. Batteries Supercaps 2, 168–177 (2019).
Soc. 151, A1977–A1988 (2004). 48. Gajan, A. et al. Solid electrolyte interphase instability in operating
28. Wang, A., Kadam, S., Li, H., Shi, S. & Qi, Y. Review on modeling of the lithium-ion batteries unraveled by enhanced-Raman spectroscopy.
anode solid electrolyte interphase (SEI) for lithium-ion batteries. npj ACS Energy Lett. 6, 1757–1763 (2021).
Comput. Mater. 4, 15 (2018). 49. Hy, S., Felix, F., Rick, J., Su, W. N. & Hwang, B. J. Direct in situ
29. Stiles, P. L., Dieringer, J. A., Shah, N. C. & Duyne, R. P. V. Surface- observation of Li2O evolution on Li-rich high-capacity cathode
enhanced Raman spectroscopy. Annu. Rev. Anal. Chem. 1, material, Li[NixLi(1-2x)/3Mn(2-x)/3]O2 (0 ≤ x ≤ 0.5). J. Am. Chem. Soc.
601–626 (2008). 136, 999–1007 (2014).
30. Wu, D.-Y., Li, J.-F., Ren, B. & Tian, Z.-Q. Electrochemical surface- 50. Qiao, Y. et al. From O2− to HO2−: reducing by-products and over-
enhanced Raman spectroscopy of nanostructures. Chem. Soc. Rev. potential in Li-O2 batteries by water addition. Angew. Chem. Int. Ed.
37, 1025–1041 (2008). 56, 4960–4964 (2017).
31. Ding, S.-Y. et al. Nanostructure-based plasmon-enhanced Raman 51. Li, C.-Y. et al. Surface changes of LiNixMnyCo1–x–yO2 in Li-ion bat-
spectroscopy for surface analysis of materials. Nat. Rev. Mater. 1, teries using in situ surface-enhanced Raman spectroscopy. J. Phys.
16021 (2016). Chem. C. 124, 4024–4031 (2020).
32. Panneerselvam, R. et al. Surface-enhanced Raman spectroscopy: 52. Martin-Yerga, D. et al. Dynamics of solid-electrolyte interphase
bottlenecks and future directions. Chem. Commun. 54, formation on silicon electrodes revealed by combinatorial electro-
10–25 (2018). chemical screening. Angew. Chem. Int. Ed. 61, e202207184 (2022).
33. Johnson, L. et al. The role of LiO2 solubility in O2 reduction in aprotic 53. Galloway, T. A., Cabo-Fernandez, L., Aldous, I. M., Braga, F. &
solvents and its consequences for Li–O2 batteries. Nat. Chem. 6, Hardwick, L. J. Shell isolated nanoparticles for enhanced Raman
1091–1099 (2014). spectroscopy studies in lithium–oxygen cells. Faraday Discuss.
34. Gogoi, N., Melin, T. & Berg, E. J. Elucidating the step-wise solid 205, 469–490 (2017).
electrolyte interphase formation in lithium-ion batteries with oper- 54. Gao, J. S. & Tian, Z. Q. Surface enhanced Raman scattering of pyr-
ando Raman spectroscopy. Adv. Mater. Interfaces 9, idine at copper electrodes excited with a 514.5 nm line. Chem. Phys.
2200945 (2022). Lett. 262, 151–154 (1996).
35. Wang, J., Zhang, Y., Guo, L., Wang, E. & Peng, Z. Identifying reactive 55. Gu, Y. et al. Designable ultra-smooth ultra-thin solid-electrolyte
sites and transport limitations of oxygen reactions in aprotic lithium- interphases of three alkali metal anodes. Nat. Commun. 9,
O2 batteries at the stage of sudden death. Angew. Chem. Int. Ed. 55, 1339 (2018).
5201–5205 (2016). 56. Gu, Y. et al. Electrochemical polishing of lithium metal surface for
36. Miroshnikov, Y. et al. Operando micro-Raman study revealing highly demanding solid-electrolyte interphase. ChemElectroChem
enhanced connectivity of plasmonic metals decorated silicon 6, 181–188 (2019).
anodes for lithium-ion batteries. ACS Appl. Energy Mater. 1, 57. Zhang, Z. et al. Capturing the swelling of solid-electrolyte inter-
1096–1105 (2018). phase in lithium metal batteries. Science 375, 66–70 (2022).

Nature Communications | (2023)14:3536 10


Article https://doi.org/10.1038/s41467-023-39192-z

58. Weber, R. et al. Long cycle life and dendrite-free lithium morphol- C.Y., and R.X. wrote the manuscript. Y.G., J.H.W., R.Y.Z, C.J.Z., G.L., H.Y.L
ogy in anode-free lithium pouch cells enabled by a dual-salt liquid and W.W.W. performed experiments. E.M.Y. conducted the COMSOL
electrolyte. Nat. Energy 4, 683–689 (2019). simulations. J.D.L. conducted the DFT calculations. S.H.L., J.L.Y., Y.Q.,
59. Louli, A. J. et al. Diagnosing and correcting anode-free cell failure J.W.Y., D.Y.W., L.Z., G.K.L. and J.F.L. helped with discussion. All authors
via electrolyte and morphological analysis. Nat. Energy 5, discussed and analyzed the data.
693–702 (2020).
60. Li, S. et al. High-efficacy and polymeric solid-electrolyte interphase Competing interests
for closely packed Li electrodeposition. Adv. Sci. 8, 2003240 (2021). The authors declare no competing interests.
61. Jiao, S. et al. Stable cycling of high-voltage lithium metal batteries
in ether electrolytes. Nat. Energy 3, 739–746 (2018). Additional information
62. Luo, S. H. et al. Developing a peak extraction and retention (PEER) Supplementary information The online version contains
algorithm for improving the temporal resolution of Raman spec- supplementary material available at
troscopy. Anal. Chem. 93, 8408–8413 (2021). https://doi.org/10.1038/s41467-023-39192-z.
63. Patnaik, P. Handbook of Inorganic Chemicals (McGraw-Hill, 2003).
64. Johnson, P. B. & Christy, R. W. Optical constants of the noble metals. Correspondence and requests for materials should be addressed to
Phys. Rev. B 6, 4370–4379 (1972). Zhong-Qun Tian, Yi Cui or Bing-Wei Mao.
65. Palik, E. D. Handbook of Optical Constants of Solid (Academic
Press, 1997). Peer review information Nature Communications thanks Guang Yang
66. Lee, C., Yang, W. & Parr, R. G. Development of the Colle-Salvetti and the other, anonymous, reviewer(s) for their contribution to the peer
correlation-energy formula into a functional of the electron density. review of this work. A peer review file is available.
Phys. Rev. B 37, 785–789 (1988).
67. Becke, A. D. Density‐functional thermochemistry. III. The role of Reprints and permissions information is available at
exact exchange. J. Chem. Phys. 98, 5648–5652 (1993). http://www.nature.com/reprints
68. Marenich, A. V., Cramer, C. J. & Truhlar, D. G. Universal solvation
model based on solute electron density and on a continuum model Publisher’s note Springer Nature remains neutral with regard to jur-
of the solvent defined by the bulk dielectric constant and atomic isdictional claims in published maps and institutional affiliations.
surface tensions. J. Phys. Chem. B 113, 6378–6396 (2009).
69. Frisch, M. J. et al. Gaussian 09 (Revision D.01) (Gaussian, Inc., Open Access This article is licensed under a Creative Commons
Wallingford, CT, 2009). Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as
Acknowledgements long as you give appropriate credit to the original author(s) and the
This work was supported by the National Natural Science Foundation of source, provide a link to the Creative Commons license, and indicate if
China (22002129 (Y.G.), 21972119 (B.W.M.), 21991151 (B.W.M.), 22202162 changes were made. The images or other third party material in this
(E.M.Y.), 22072123 (J.W.Y.), 22102137 (W.W.W.)), the China Postdoctoral article are included in the article’s Creative Commons license, unless
Science Foundation (2019TQ0177 (Y.G.), 2022M722648 (E.M.Y.), indicated otherwise in a credit line to the material. If material is not
2022T150548 (W.W.W.)). We thank H.X. Lin, C.Y. Li, and F.R. Fan for included in the article’s Creative Commons license and your intended
helpful discussions and advice, Y.H. Hong for experimental assistance use is not permitted by statutory regulation or exceeds the permitted
and C. Geng for help on schematic design. use, you will need to obtain permission directly from the copyright
holder. To view a copy of this license, visit http://creativecommons.org/
Author contributions licenses/by/4.0/.
Z.Q.T., B.W.M., and Y.G. conceived idea. Y.G., B.W.M., and Z.Q.T.
designed experiments and analyzed results. Y.G., E.M.Y., B.W.M., Z.Q.T, © The Author(s) 2023

Nature Communications | (2023)14:3536 11

You might also like